Standardized pseudospectral formulation of the inviscid...

28
Standardized pseudospectral formulation of the inviscid supsersonic blunt body problem q Gregory P. Brooks a,1 , Joseph M. Powers b, * ,2 a Air Force Research Laboratory, Wright-Patterson AFB, OH 45433, USA b Department of Aerospace and Mechanical Engineering, University of Notre Dame, 372 Fitzpatrick Hall of Engineering, Notre Dame, IN 46556-5637, USA Received 21 February 2003; received in revised form 24 November 2003; accepted 25 November 2003 Available online 23 December 2003 Abstract A highly accurate pseudospectral numerical approximation to the generalized coordinate, nonconservative form of the Euler equations is implemented for supersonic flow over an axisymmetric blunt body geometry; shock fitting is employed to maintain global accuracy and minimize the corrupting influence of numerical viscosity. The variables in the Euler equations as well as the physical grid coordinates are collocated via Lagrange interpolating polynomials and the problem is then cast in the standard form of a large system of ordinary differential equations, dx=ds ¼ qðxÞ, which can be solved using standard solution techniques that do not require an explicit criteria for the minimum time step. Code verification is performed by demonstrating through a series of grid refinement tests that the error in the ap- proximation to a Taylor–Maccoll solution converges to 10 12 . Grid refinement tests for flow over a blunt body show convergence of the numerical error also to 10 12 . The code is validated for supersonic flow over a blunt body by comparison with the modified Newtonian approximation for the surface pressure distribution and empirical predictions for the shock shape. The ability of the method to capture unsteady flow phenomena is demonstrated on the problem of a planar acoustic wave interacting with an attached shock. Ó 2003 Elsevier Inc. All rights reserved. Keywords: Shock fitting; Supersonic blunt body; Euler equations; Pseudospectral method; System of ordinary differential equations Journal of Computational Physics 197 (2004) 58–85 www.elsevier.com/locate/jcp q This study has received primary support from the United States Air Force Palace Knight Program and additional support from Los Alamos National Laboratory. * Corresponding author. Tel.: +574-631-5978; fax: +574-631-8341. E-mail addresses: [email protected] (G.P. Brooks), [email protected] (J.M. Powers). 1 Research Aerospace Engineer. 2 Associate Professor. 0021-9991/$ - see front matter Ó 2003 Elsevier Inc. All rights reserved. doi:10.1016/j.jcp.2003.11.017

Transcript of Standardized pseudospectral formulation of the inviscid...

  • Journal of Computational Physics 197 (2004) 58–85

    www.elsevier.com/locate/jcp

    Standardized pseudospectral formulation of the inviscidsupsersonic blunt body problem q

    Gregory P. Brooks a,1, Joseph M. Powers b,*,2

    a Air Force Research Laboratory, Wright-Patterson AFB, OH 45433, USAb Department of Aerospace and Mechanical Engineering, University of Notre Dame, 372 Fitzpatrick Hall of Engineering,

    Notre Dame, IN 46556-5637, USA

    Received 21 February 2003; received in revised form 24 November 2003; accepted 25 November 2003

    Available online 23 December 2003

    Abstract

    A highly accurate pseudospectral numerical approximation to the generalized coordinate, nonconservative form of

    the Euler equations is implemented for supersonic flow over an axisymmetric blunt body geometry; shock fitting is

    employed to maintain global accuracy and minimize the corrupting influence of numerical viscosity. The variables in

    the Euler equations as well as the physical grid coordinates are collocated via Lagrange interpolating polynomials and

    the problem is then cast in the standard form of a large system of ordinary differential equations, dx=ds ¼ qðxÞ, whichcan be solved using standard solution techniques that do not require an explicit criteria for the minimum time step.

    Code verification is performed by demonstrating through a series of grid refinement tests that the error in the ap-

    proximation to a Taylor–Maccoll solution converges to 10�12. Grid refinement tests for flow over a blunt body show

    convergence of the numerical error also to 10�12. The code is validated for supersonic flow over a blunt body by

    comparison with the modified Newtonian approximation for the surface pressure distribution and empirical predictions

    for the shock shape. The ability of the method to capture unsteady flow phenomena is demonstrated on the problem of

    a planar acoustic wave interacting with an attached shock.

    � 2003 Elsevier Inc. All rights reserved.

    Keywords: Shock fitting; Supersonic blunt body; Euler equations; Pseudospectral method; System of ordinary differential equations

    qThis study has received primary support from the United States Air Force Palace Knight Program and additional support from Los

    Alamos National Laboratory.*Corresponding author. Tel.: +574-631-5978; fax: +574-631-8341.

    E-mail addresses: [email protected] (G.P. Brooks), [email protected] (J.M. Powers).1 Research Aerospace Engineer.2 Associate Professor.

    0021-9991/$ - see front matter � 2003 Elsevier Inc. All rights reserved.doi:10.1016/j.jcp.2003.11.017

    mail to: [email protected]

  • G.P. Brooks, J.M. Powers / Journal of Computational Physics 197 (2004) 58–85 59

    1. Introduction

    As a prerequisite for developing a new technique for shape optimization of bodies in supersonic flow[1], we require a highly accurate flow solution technique. Consequently, in this work, we present a

    pseudospectral numerical approximation to the solution for the supersonic, inviscid flow of a calorically

    perfect ideal gas over an axisymmetric blunt body in standardized form. The shock is fitted since ap-

    proximation of discontinuous solutions with high order polynomials exhibits the Gibbs phenomenon in

    the form of global oscillations in the solution [2]. These oscillations can cause the numerical scheme to

    become unstable, and attempts to remove the oscillations by spectral filtering or by addition of artificial

    viscosity significantly reduces the accuracy of the numerical method. The more common alternative of

    shock capturing, while generally stable and nonoscillatory, yields only first order accuracy [3].While the main motivation of our work is to have high accuracy steady state solutions available for use

    in a shape optimization procedure, we note there are outstanding questions regarding numerical and

    physical instabilities in multi-dimensional shock dynamics for which a standard formulation of

    dx=ds ¼ qðxÞ has value. As stated in [4–11], the issues of shock stability are controversial since it is not clearwhat effects are real and which ones are purely numerical artifacts. Although shock stability will not be

    investigated in the current work, the shock fitting method used here is the most appropriate for addressing

    fundamental stability questions for inviscid flows due to the significantly low level of numerical viscosity

    inherent in the method. Furthermore, should a physical instability exist, one would like to have a numericalmethod with sufficient robustness to automatically select the time step so as to resolve the dynamics, and

    because of the additional dynamic equations which arise from the unsteady shock wave, a simple CFL

    criterion is insufficient in the general case. Fortunately, the casting of the discretized governing equations

    into the form dx=ds ¼ qðxÞ allows the use of standard software packages which have automatic time stepselection, precluding the necessity of an explicit condition for numerical stability. As such, we place em-

    phasis here on obtaining this standard form.

    We briefly review the literature on solutions to the supersonic flow over blunt body geometries.

    Rusanov [12] and Hayes and Probstein [13] have given thorough reviews of early contributions, ofwhich a few will be mentioned. Two methodologies for calculating solutions to the supersonic flow

    about a blunt body are the direct and inverse methods. In the direct method the body shape is

    specified, and then the shock shape and flow field are calculated. In the inverse method the shock shape

    is specified, and the body shape which would support that shock shape is calculated. At first, studies

    concerning the inverse problem were based on series expansion of the governing equations in the vi-

    cinity of the shock wave [14]. Later, numerical solutions to the inverse, supersonic blunt body problem

    were performed by Garabedian and Lieberstein [15] and Van Dyke [16]. Evans and Harlow [17] were

    the first to generate numerical solutions to the direct problem by integrating the unsteady Eulerequations to a relaxed steady state solution. Moretti and Abbett [18] used finite differences and fitting

    of the shock to generate accurate solutions of the Euler equations about a blunt body; this laid the

    foundation for subsequent shock fitting numerical schemes. Pseudospectral approximations to the Euler

    equations employing shock fitting with spectral filtering were first performed by Hussaini et al. [19] and

    without spectral filtering by Kopriva [20]. The numerical technique employed in the current paper

    builds on the work by Kopriva [20,21], and Brooks and Powers [22].

    We briefly review the pseudospectral methods. An early unified mathematical description of the theory

    of spectral and pseudospectral methods was given by Gottlieb and Orszag [2]. Significant advances oc-curred in the late 1970s and early 1980s and are well documented by Canuto et al. [23], with particular

    application to fluid dynamics. For a more recent review, see [24].

    There does not appear to be complete consensus in the literature for the definition of pseudospectral; a

    definition is adopted here which we believe useful and consistent with that of Fornberg [25]. We define a

    pseudospectral method to be a collocation type of method of weighted residuals, as defined by Finlayson

  • 60 G.P. Brooks, J.M. Powers / Journal of Computational Physics 197 (2004) 58–85

    [26], in which the error in the solution to the governing equations is driven to zero at collocation points; the

    flow quantities are represented in terms of global Lagrange interpolating polynomials defined at the col-

    location points. The spatial derivatives of the flow quantities are then calculated by differentiating theLagrange interpolating polynomials. Efficient algorithms for calculating derivatives of Lagrange interpo-

    lating polynomials on arbitrary grids can be found in [25]; these algorithms were used in the current work.

    A property of the pseudospectral method is that approximations to derivatives have global support, making

    it equivalent to a finite difference scheme with a stencil that extends over the entire domain. As the number

    of points is increased, the size of the stencil grows, leading to a higher order accurate solution.

    2. Supersonic blunt body flow and pseudospectral solver

    2.1. Standard formulation

    Before presenting the geometry, governing equations, boundary and initial conditions for the blunt body

    problem, we will briefly outline the procedure for formulating this problem as a system of ordinary dif-

    ferential equations (ODEs). The Euler equations, physical grid evolution equations, shock velocity equa-

    tion, and boundary conditions defined over the computational domain

    X : n 2 ½0; 1�; g 2 ½0; 1�f g; ð1Þ

    and bounded by S, can be written in the form of the following coupled system of time-dependent partialdifferential and algebraic equations

    oy

    osþ f y; oy

    on;oy

    og

    � �¼ 0; ð2Þ

    g y;oy

    on;oy

    og

    � �¼ 0; ð3Þ

    along with the initial conditions

    y n; g; 0ð Þ ¼ y0 n; gð Þ; ð4Þ

    where n and g are independent spatial variables in the computational space, and f and g are nonlinearfunctions of the dependent variables yðn; g; sÞ and its spatial derivatives. All of the algebraic constraints,Eq. (3), are boundary conditions and thus apply only on S.

    Approximating yðn; g; sÞ and its spatial derivatives via global Lagrange interpolating polynomials, thesystem of partial differential and algebraic equations in Eqs. (2) and (3) reduce to the following system of P2differential algebraic equations:

    dypðsÞds

    ¼ fp y1; . . . ; yP2ð Þ; p ¼ 1; . . . ; P1; ð5Þ

    0 ¼ gp0 y1; . . . ; yP2ð Þ; p0 ¼ P1 þ 1; . . . ; P2; ð6Þ

    with initial conditions from Eq. (4)

    ypð0Þ ¼ y0p; p ¼ 1; . . . ; P1: ð7Þ

  • G.P. Brooks, J.M. Powers / Journal of Computational Physics 197 (2004) 58–85 61

    Here y1ðsÞ; . . . ; yP2ðsÞ are the flow quantities, physical grid coordinates, and shock velocity evaluated atnodal points on an ðN þ 1Þ � ðM þ 1Þ mesh. Solving for the yp0 ðsÞ, p0 ¼ P1 þ 1; . . . ; P2, as explicit functionsof the yp, p ¼ 1; . . . ; P1, i.e.

    yp0 ðsÞ ¼ bgp0 y1; . . . ; yP1ð Þ; p0 ¼ P1 þ 1; . . . ; P2: ð8ÞEqs. (5) and (6) are converted into the following system of P1 ODEs:

    dxpðsÞds

    ¼ qp x1; . . . ; xP1ð Þ; p ¼ 1; . . . ; P1; ð9Þ

    where

    qp x1; . . . ; xP1ð Þ � fp y1; . . . ; yP1 ; bgp0 y1; . . . ; yP1ð Þ� �; p ¼ 1; . . . ; P1; p0 ¼ P1 þ 1; . . . ; P2; ð10Þand

    xpðsÞ ¼ ypðsÞ; p ¼ 1; . . . ; P1: ð11Þ

    The accompanying initial conditions are

    xpð0Þ ¼ x0p; p ¼ 1; . . . ; P1: ð12Þ

    In compact vector notation, Eq. (9) is

    dx

    ds¼ qðxÞ; ð13Þ

    with accompanying initial conditions

    xð0Þ ¼ x0 ð14Þ

    from Eq. (12).

    2.2. Governing equations

    The two-dimensional, axisymmetric Euler equations for a calorically perfect ideal gas are in dimen-

    sionless form:

    oqot

    þ u oqor

    þ w oqoz

    þ q ouor

    �þ ow

    ozþ u

    r

    �¼ 0; ð15Þ

    ouot

    þ u ouor

    þ w ouoz

    þ 1qopor

    ¼ 0; ð16Þ

    owot

    þ u owor

    þ w owoz

    þ 1qopoz

    ¼ 0; ð17Þ

    opot

    þ u opor

    þ w opoz

    þ cp ouor

    �þ ow

    ozþ u

    r

    �¼ 0; ð18Þ

  • 62 G.P. Brooks, J.M. Powers / Journal of Computational Physics 197 (2004) 58–85

    where q is density, p is pressure, u and w are the velocities in the radial and axial directions, respectively, r isthe radial coordinate, z is the axial coordinate, t is time, and c is the ratio of specific heats. The dimensionalform for pressure, p�, density, q�, and r� and z� components of velocity, u� and w�, respectively, are re-covered by the following equations:

    p� ¼ pp�1; ð19Þ

    q� ¼ qq�1; ð20Þ

    u� ¼ uffiffiffiffiffiffiffiffiffiffiffiffiffiffiffip�1=q

    �1

    q; w� ¼ w

    ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffip�1=q

    �1

    q; ð21Þ

    where dimensional quantities are denoted by a �, and freestream quantities are denoted by 1. The di-mensional space and time variables are

    z� ¼ zL; r� ¼ rL; ð22Þ

    t� ¼ tL=ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffip�1=q

    �1

    q; ð23Þ

    where L is the length of the body. The freestream flow is at zero angle of attack so that the component offreestream velocity in the r direction, u1 ¼ 0.

    Defining the entropy to be s, we have, for a calorically perfect ideal gas with zero freestream entropy,

    s ¼ ln pqc

    � �; ð24Þ

    where the entropy is nondimensionalized by the the specific heat at constant volume, c�v,

    s� ¼ sc�v: ð25Þ

    To facilitate the solution to the Euler equations for time-varying geometry, Eqs. (15)–(18) are rewritten

    in terms of a general body-fitted coordinate system, nðz; r; tÞ, gðz; r; tÞ and sðz; r; tÞ. Employing the chain ruleof differentiation

    o

    oz¼ on

    ozo

    onþ og

    ozo

    ogþ os

    ozo

    os;

    o

    or¼ on

    oro

    onþ og

    oro

    ogþ os

    oro

    os;

    o

    ot¼ on

    oto

    onþ og

    oto

    ogþ os

    oto

    os;

    ð26Þ

    and taking sðz; r; tÞ ¼ t, the nondimensional form of Eqs. (15)–(18) in generalized coordinates is

    oqos

    þ bu oqon

    þ bw oqog

    þ q onor

    ouon

    �þ on

    ozowon

    þ ogor

    ouog

    þ ogoz

    owog

    �þ qu

    r¼ 0; ð27Þ

    ouos

    þ bu ouon

    þ bw ouog

    þ 1q

    onor

    opon

    �þ og

    oropog

    �¼ 0; ð28Þ

  • G.P. Brooks, J.M. Powers / Journal of Computational Physics 197 (2004) 58–85 63

    owos

    þ bu owon

    þ bw owog

    þ 1q

    onoz

    opon

    �þ og

    ozopog

    �¼ 0; ð29Þ

    opos

    þ bu opon

    þ bw opog

    þ cp onor

    ouon

    �þ on

    ozowon

    þ ogor

    ouog

    þ ogoz

    owog

    �þ cpu

    r¼ 0; ð30Þ

    where the contravariant velocity components bu and bw arebu ¼ on

    otþ u on

    orþ w on

    oz; ð31Þ

    bw ¼ ogot

    þ u ogor

    þ w ogoz

    : ð32Þ

    The following standard relations between the metrics and inverse metrics will be necessary

    onoz

    ¼ 1Jorog

    ;ogoz

    ¼ � 1Joron

    ;

    onor

    ¼ � 1Jozog

    ;ogor

    ¼ 1Jozon

    ;

    onot

    ¼oros

    ozog � orog ozos

    � �J

    ;ogot

    ¼oron

    ozos � oros ozon

    � �J

    ;

    J ¼ orog

    ozon

    � oron

    ozog

    ;

    ð33Þ

    where J is the determinant of the metric Jacobian matrix.

    −0.4 −0.2 0.2 0.4 0.6 0.8 1

    0.6

    0.8

    1

    1.2

    0.2

    0.4

    z

    r

    η

    ξ Shock

    v∞

    h(ξ,τ)

    Body (R=Z b)

    Fig. 1. Schematic of shock-fitted high Mach number flow over an axisymmetric blunt body including computational ðn; gÞ andphysical ðr; zÞ coordinates.

  • 64 G.P. Brooks, J.M. Powers / Journal of Computational Physics 197 (2004) 58–85

    2.3. Computational and physical coordinates

    The physical domain of the blunt body problem, Fig. 1, is mapped to the computational domain,n 2 ½0; 1�, g 2 ½0; 1�, in such a way that the body surface lies along the computational boundary n; 0ð Þ, theshock lies along the boundary ðn; 1Þ, the symmetry axis is a third boundary at ð0; gÞ, and the fourthboundary at ð1; gÞ is a supersonic outflow. The transformation between the physical coordinates ðr; zÞ andcomputational coordinates n; gð Þ is taken to be

    r n; g; sð Þ ¼ RðnÞ þg dZðnÞ

    dn h n; sð ÞffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffidRðnÞdn

    � �2þ dZðnÞ

    dn

    � �2r ; ð34Þ

    z n; g; sð Þ ¼ ZðnÞ �g dRðnÞ

    dn h n; sð ÞffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffidRðnÞdn

    � �2þ dZðnÞ

    dn

    � �2r ; ð35Þ

    where the nonlinear function hðn; sÞ must be specified to completely determine the mapping, and RðnÞ andZðnÞ are known functions. After manipulation, the transformations in Eqs. (34) and (35) yield the followingidentity:

    h n; sð Þ ¼ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiz n; 1; sð Þ � z n; 0; sð Þð Þ2 þ r n; 1; sð Þ � r n; 0; sð Þð Þ2

    q; ð36Þ

    from which it is seen that the function hðn; sÞ is the distance in r–z space between the body surface, g ¼ 0,and the shock, g ¼ 1, along lines of constant n. The function hðn; sÞ is subsequently referred to as the shockdistance function. We see that Eqs. (34) and (35) form an implicit algebraic equation for the coordinate

    transformation. It is apparent from Eqs. (34) and (35) that the functions RðnÞ and ZðnÞ parameterize theblunt body surface, g ¼ 0, i.e.

    r n; 0; sð Þ ¼ RðnÞ;z n; 0; sð Þ ¼ ZðnÞ;

    ð37Þ

    and that the body surface is not a function of time. The transformations in Eqs. (34) and (35) have been

    constructed so that lines of constant n are normal to the body surface and have no curvature in r–z space.The time evolution equations for the physical grid rðn; g; sÞ, and zðn; g; sÞ can be found by differentiating

    Eqs. (34) and (35) with respect to time as follows:

    o

    osr n; g; sð Þ ¼

    g dZðnÞdn v n; sð Þffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

    dRðnÞdn

    � �2þ dZ nð Þ

    dn

    � �2r ; ð38Þ

    o

    osz n; g; sð Þ ¼ �

    g dRðnÞdn v n; sð Þffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

    dRðnÞdn

    � �2þ dZ nð Þ

    dn

    � �2r ; ð39Þwhere the shock velocity function v n; sð Þ is

    v n; sð Þ ¼ oos

    h n; sð Þ: ð40Þ

  • G.P. Brooks, J.M. Powers / Journal of Computational Physics 197 (2004) 58–85 65

    2.4. Boundary conditions

    The kinematic boundary condition of no mass flux at the body surface requires that the velocity com-

    ponent normal to the body surface, vBNðn; sÞ, be equal to zero, i.e.

    vBN n; sð Þ ¼ bw n; 0; sð Þ ¼ 0: ð41ÞIn order to formulate a numerical boundary condition at the body for q, u, w, and p, Eqs. (27)–(30) are

    written in the following form:

    oz

    osþ A oz

    onþ B oz

    ogþ s ¼ 0; ð42Þ

    where

    z ¼

    quwp

    26643775; s ¼

    qu=r0

    0

    cpu=r

    26643775; ð43Þ

    A ¼

    bu q onor q

    onoz 0

    0 bu 0 1q onor0 0 bu 1q onoz0 cp on

    or cponoz bu

    2666664

    3777775; B ¼bw q og

    or qogoz 0

    0 bw 0 1q ogor0 0 bw 1q ogoz0 cp og

    or cpogoz bw

    2666664

    3777775: ð44Þ

    The flux Jacobian matrix B is then decomposed as

    B ¼ P�1KgP; ð45Þ

    where the square matrix P contains the left eigenvectors of B in its rows; the diagonal matrix Kg containsthe eigenvalues of B in its diagonal; and P�1 is the inverse of P. Substituting Eq. (45) into Eq. (42) and

    premultiplying by P yields the following characteristic formulation [27,28] of the governing equations:

    Poz

    osþ PA oz

    onþ KgP

    oz

    ogþ Ps ¼ 0: ð46Þ

    The diagonal eigenvalue matrix Kg and the left eigenvector matrix P are

    Kg ¼

    bw 0 0 00 bw 0 00 0 bw � c ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiog

    oz

    � �2 þ ogor

    � �2q0

    0 0 0 bw þ c ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiogoz

    � �2 þ ogor

    � �2q

    26666664

    37777775; ð47Þ

  • 66 G.P. Brooks, J.M. Powers / Journal of Computational Physics 197 (2004) 58–85

    P ¼

    0 �ogozogor

    ogozð Þ

    2þ og

    orð Þ2

    ogozð Þ

    2

    ogozð Þ

    2þ og

    orð Þ2 0

    1 0 0 � 1c20 � qc

    ogoz

    2

    ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiogozð Þ

    2þ og

    orð Þ2

    q � qcogor2

    ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiogozð Þ

    2þ og

    orð Þ2

    q 12

    0qcog

    oz

    2

    ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiogozð Þ

    2þ og

    orð Þ2

    q qcogor2

    ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiogozð Þ

    2þ og

    orð Þ2

    q 12

    266666666664

    377777777775; ð48Þ

    where c ¼ffiffiffiffiffiffiffiffiffifficp=q

    pis the dimensionless acoustic speed. Since bw is everywhere negative and since

    kbwk < kc ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiðog=ozÞ2 þ ðog=orÞ2q k at g ¼ 0, only the first three of the equations in Eq. (46) can be used informulating numerical boundary conditions since they are associated with negative eigenvalues, Kg. Thefourth equation in Eq. (46) is associated with a positive eigenvalue and thus describes information prop-

    agation from inside the body which must therefore be discarded as nonphysical; in its place the physical

    boundary condition, Eq. (41), is employed. Making use of the fact that bwðn; 0; sÞ ¼ obwos jðn;0;sÞ ¼ 0, the three

    differential equations from Eq. (46) to be solved at the body surface, g ¼ 0, are cast in the following form,which is consistent with Eq. (2)

    oqos

    �¼ 1

    c2opos

    � bu oqon

    �� 1c2

    opon

    n;0;sð Þ

    ; ð49Þ

    o

    osvBT n; sð Þ ¼

    bu ogor

    owon �

    ogoz

    ouon

    � �þ 1q

    ogor

    onoz �

    ogoz

    onor

    � �opon

    � �ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiogoz

    � �2 þ ogor

    � �2q

    n;0;sð Þ

    ; ð50Þ

    opos

    264 ¼ qcbu ogor ouon þ ogoz owon� �

    þ c ogor

    onor þ

    ogoz

    onoz

    � �oponffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

    ogoz

    � �2 þ ogor

    � �2q � qc2 onor ouon�

    þ onoz

    owon

    þ ogor

    ouog

    þ ogoz

    owog

    þ c

    ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiogoz

    � �2þ og

    or

    � �2sopog

    � bu opon

    � qc2ur

    375

    n;0;sð Þ

    : ð51Þ

    The velocity components uðn; 0; sÞ and wðn; 0; sÞ are specified as following functions of vBT:

    u n; 0; sð Þ ¼oron vBT n; sð Þffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffioron

    � �2þ oz

    on

    � �2r

    n;0;sð Þ

    ; w n; 0; sð Þ ¼ozon vBT n; sð Þffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffioron

    � �2þ oz

    on

    � �2r

    n;0;sð Þ

    : ð52Þ

    At the shock boundary, the Rankine–Hugoniot relations are solved along with a compatibility equation.

    Specifically, the Rankine–Hugoniot relations are

    v1 � eSTj n;1;sð Þ ¼ v � eSTj n;1;sð Þ; ð53Þ

    dS n; sð Þ ¼c� 1cþ 1 d1 n; sð Þ þ

    2cp1cþ 1ð Þq1d1 n; sð Þ

    ; ð54Þ

  • G.P. Brooks, J.M. Powers / Journal of Computational Physics 197 (2004) 58–85 67

    p n; 1; sð Þ ¼ 2q1cþ 1 d

    21 n; sð Þ �

    c� 1cþ 1 p1; ð55Þ

    q n; 1; sð Þ ¼ d1 n; sð ÞdS n; sð Þ

    q1; ð56Þ

    where dS and d1 are the component of fluid velocity normal to the shock in the shock-attached referenceframe on the downstream and freestream sides of the shock, respectively, i.e.

    dS n; sð Þ ¼v � eSNj n;1;sð Þ � vSN n; sð Þ;v � eSNj n;1;sð Þ � eg � eSN

    � �v n; sð Þ;

    �ð57Þ

    d1 n; sð Þ ¼ v1 � eSNj n;1;sð Þ � eg � eSN� �

    v n; sð Þ; ð58Þ

    and vðn; g; sÞ is the nondimensional velocity vector. The nondimensional freestream velocity vector v1 is

    v1 ¼ffiffiffiffiffiffiffifficp1q1

    rM1ez: ð59Þ

    In Eqs. (53)–(57), eST is a unit vector in the direction tangent to the shock wave, eSN is a unit vector in thedirection normal to the shock wave, and vSNðn; sÞ and vðn; sÞ are the velocities of the shock in the eSN and egdirections, respectively. Quantities denoted with a subscript of 1 are freestream quantities, and those withno subscripts are post-shock quantities. The unit vectors eST and eSN are in terms of the inverse metrics

    eST ¼� og

    oz er þogor ezffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

    ogoz

    � �2 þ ogor

    � �2q

    n;1;sð Þ

    ; ð60Þ

    eSN ¼ogor er þ

    ogoz ezffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

    ogoz

    � �2 þ ogor

    � �2q

    n;1;sð Þ

    : ð61Þ

    In order to solve the Rankine–Hugoniot equations, an expression for the shock velocity, vðn; sÞ, isneeded. Differentiating Eqs. (54) and (55) with respect to time yields

    o

    osdS n; sð Þ ¼ A1 n; sð Þ

    o

    osd1 n; sð Þ; ð62Þ

    o

    osp n; 1; sð Þ ¼ A2 n; sð Þ

    o

    osd1 n; sð Þ; ð63Þ

    where

    A1 n; sð Þ ¼c� 1cþ 1�

    2c

    cþ 1ð Þd21 n; sð Þ; A2 n; sð Þ ¼

    4d1 n; sð Þcþ 1 : ð64Þ

    The terms ðo=osÞdSðn; sÞ and ðo=osÞd1ðn; sÞ in Eqs. (62) and (63) are found by differentiating Eqs. (57) and(58), respectively, to yield the following:

  • 68 G.P. Brooks, J.M. Powers / Journal of Computational Physics 197 (2004) 58–85

    o

    osdS n; sð Þ ¼

    ov

    os� eSN

    �þ v � oeSN

    os� eg � eSN� � ov

    os� veg �

    oeSN

    os

    n;1;sð Þ

    ; ð65Þ

    o

    osd1 n; sð Þ ¼

    ov1

    os� eSN

    �þ v1 �

    oeSN

    os� eg � eSN� � ov

    os� veg �

    oeSN

    os

    n;1;sð Þ

    : ð66Þ

    Multiplying Eq. (62) by qc and adding it to Eq. (63), and replacing the term ðo=osÞdSðn; sÞ by Eq. (65)and the term ðo=osÞd1ðn; sÞ by (66), we arrive at the following equation for the shock accelerationðo=osÞvðn; sÞ,

    o

    osv n; sð Þ ¼

    A2 þ qcA1ð Þ v1 � veg� �

    � qc v� veg� ��

    � oeSNos � qc ovos � eSN �

    opos

    eg � eSN� �

    A2 þ qc A1 � 1ð Þ½ �

    n;1;sð Þ

    : ð67Þ

    The terms op=os and qcðov=osÞ � eSN must be specified by a compatibility equation which is the charac-teristic equation associated with the wave propagating from the body to the shock along the normal di-

    rection. This compatibility equation is in the same form as the fourth compatibility equation in Eq. (46)

    only written in shock coordinates instead of the body coordinate system ðn; g; sÞ. After some simplification,the following shock acceleration equation is obtained:

    o

    osv n; sð Þ ¼

    A2 þ qcA1ð Þ v1 � veg� �

    � oeSNos � qc v� veg

    � �� oeSN

    os þ A3eg � eSN� �

    A2 þ qc A1 � 1ð Þ½ �

    n;1;sð Þ

    ; ð68Þ

    where

    A3 ¼ bu opon

    þ bw opog

    þ cp onoz

    owon

    �þ og

    ozowog

    þ onor

    ouon

    þ ogor

    ouog

    þ qcozonffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

    ozon

    � �2þ or

    on

    � �2r bu ouron�2664 þ bw ouogþ 1q onor opon

    �þ og

    oropog

    ��

    �oronffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

    ozon

    � �2þ or

    on

    � �2r bu owon�

    þ bw owog

    þ 1q

    onoz

    opon

    �þ og

    ozopog

    ��3775þ cpur : ð69Þ

    The time derivative of the normal unit vector, oeSN=os, is found by taking the time derivative of Eq. (61)with the metrics from Eq. (33) in place of the inverse metrics to yield

    oeSN

    os¼

    oron

    o2zos on � ozon o

    2ros on

    � �ozon ez þ oron er

    � �ozon

    � �2þ or

    on

    � �2� �3=2 : ð70ÞSince there is a geometric singularity in Eq. (69) at r ¼ 0, the following alternate expression for vð0; sÞ isemployed

  • G.P. Brooks, J.M. Powers / Journal of Computational Physics 197 (2004) 58–85 69

    o

    onv 0; sð Þ ¼ 0: ð71Þ

    We impose the following appropriate boundary conditions on the centerline, n ¼ 0, in computationalcoordinates

    owon

    0;g;sð Þ

    ¼ 0; ð72Þ

    opon

    0;g;sð Þ

    ¼ 0; ð73Þ

    u 0; g; sð Þ ¼ 0; ð74Þ

    q 0; g; sð Þ ¼ q 0; 1; sð Þ p 0; g; sð Þp 0; 1; sð Þ

    � �1=c: ð75Þ

    The boundary condition in Eq. (75) comes from casting the energy equation, Eq. (18), in terms of the

    nondimensional entropy, s,

    osot

    þ u osor

    þ w osoz

    ¼ 0: ð76Þ

    Enforcing steady state, os=ot ¼ 0, and zero velocity in the r direction, uð0; g; sÞ ¼ 0, Eq. (76) reduces toosoz jð0;g;sÞ ¼ 0. Thus sð0; g; sÞ is constant and equal to sð0; 1; sÞ, the nondimensional value of the entropydownstream of the shock. Substituting sð0; g; sÞ ¼ lnðpð0; 1; sÞ=qð0; 1; sÞcÞ into the equation for entropy, Eq.(24), and simplifying gives the boundary condition in Eq. (75). We note that the enforcement of steady statefor entropy is artificial and potentially precludes some classes of unsteady behavior.

    At the supersonic outflow boundary, n ¼ 1, no physical boundary conditions are required as all wavesare exiting the domain. Here the governing equations are solved in exactly the same manner as in the

    interior.

    2.5. Summary of governing equations and boundary conditions

    The governing equations and boundary conditions can be written in terms of the system of time-de-pendent partial differential and algebraic equations in Eqs. (2)–(4) in the two space dimensions, n and g,where

    y n; g; sð Þ ¼

    q n; g; sð Þu n; g; sð Þw n; g; sð Þp n; g; sð Þr n; g; sð Þz n; g; sð ÞvBT n; sð Þq n; 0; sð Þp n; 0; sð Þv n; sð Þ

    2666666666666664

    3777777777777775; ð77Þ

  • 70 G.P. Brooks, J.M. Powers / Journal of Computational Physics 197 (2004) 58–85

    f ¼

    �bu oqon

    � bw oqog

    � q onor

    ouon

    þ onoz

    owon

    þ ogor

    ouog

    þ ogoz

    owog

    � �� qu

    rEq:ð27Þ

    �bu ouon

    � bw ouog

    � 1q

    onor

    opon

    þ ogor

    opog

    � �Eq:ð28Þ

    �bu owon

    � bw owog

    � 1q

    onoz

    opon

    þ ogoz

    opog

    � �Eq:ð29Þ

    �bu opon

    � bw opog

    � cp onor

    ouon

    þ onoz

    owon

    þ ogor

    ouog

    þ ogoz

    owog

    � �� cpu

    rEq:ð30Þ

    goZ nð Þon

    v n; sð ÞffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffioR nð Þon

    � �2þ oZ nð Þ

    on

    � �2s Eq:ð38Þ

    �goRðnÞon

    v n; sð ÞffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffioR nð Þon

    � �2þ

    oZ nð Þon

    � �2s Eq:ð39Þ

    �bu og

    orowon

    � ogoz

    ouon

    � �þ 1q

    ogor

    onoz

    � ogoz

    onor

    � �opon

    � ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiogoz

    � �2þ

    ogor

    � �2s

    n;0;sð Þ

    Eq:ð50Þ

    � 1c2opos

    þ bu oqon

    � 1c2

    opon

    � ��

    n;0;sð Þ

    Eq:ð49Þ

    �qcbu og

    orouon

    þ ogoz

    owon

    � �þ c og

    oronor

    þ ogoz

    onoz

    � �oponffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

    ogoz

    � �2þ og

    or

    � �2s266664

    �qc2 onor

    ouon

    þ onoz

    owon

    þ ogor

    ouog

    þ ogoz

    owog

    � �

    þc

    ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiogoz

    � �2þ og

    or

    � �2sopog

    � bu opon

    � qc2ur

    #

    n;0;sð Þ

    Eq:ð51Þ

    A2 þ qcA1ð Þ v1 � veg� �

    � oeSNos

    � qc v� veg� �

    � oeSNos

    þ A3eg � eSN� �

    A2 þ qc A1 � 1ð Þ½ �

    264375

    n;1;sð Þ

    Eq:ð68Þ

    26666666666666666666666666666666666666666666666666666666666666666666666666666666666666666664

    37777777777777777777777777777777777777777777777777777777777777777777777777777777777777777775

    ; ð78Þ

  • G.P. Brooks, J.M. Powers / Journal of Computational Physics 197 (2004) 58–85 71

    g ¼

    u�vBT oronffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

    oron

    � �2þ oz

    on

    � �2s266664

    377775

    n;0;sð Þ

    w�vBT ozonffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

    oron

    � �2þ oz

    on

    � �2s266664

    377775

    n;0;sð Þ

    9>>>>>>>>>>>>>>>>=>>>>>>>>>>>>>>>>;

    Eq:ð52Þ

    q n; 1; sð Þ � d1 n; sð ÞdS n; sð Þ

    Eq:ð56Þ

    v1 � eST � v � eST½ �j n;1;sð Þ Eq:ð53Þ

    dS n; sð Þ �c� 1cþ 1 d1 n; sð Þ þ

    2cp1cþ 1ð Þq1d1 n; sð Þ

    Eq:ð54Þ

    p n; 1; sð Þ � 2q1cþ 1 d

    21 n; sð Þ �

    c� 1cþ 1 p1 Eq:ð55Þ

    q 0; g; sð Þ � q 0; 1; sð Þ p 0; g; sð Þp 0; 1; sð Þ

    � �1=cEq:ð75Þ

    u 0; g; sð Þ Eq:ð74Þowon

    0;g;sð Þ

    Eq:ð72Þ

    opon

    0;g;sð Þ

    Eq:ð73Þ

    ovon

    0;sð Þ

    Eq:ð71Þ

    266666666666666666666666666666666666666666666666666664

    377777777777777777777777777777777777777777777777777775

    : ð79Þ

    0.2 0.4 0.6 0.8 1

    0.2

    0.4

    0.6

    0.8

    1

    ξ

    η

    Body

    Outflow

    Shock

    Centerline

    Fig. 2. Gauss–Lobatto Chebyshev computational grid for the shock-fitted blunt body.

  • 72 G.P. Brooks, J.M. Powers / Journal of Computational Physics 197 (2004) 58–85

    In Eqs. (77)–(79) yðn; g; sÞ : R3 ! R10 and f : R3 ! R10 while g : R3 ! R11. The functions yðn; g; sÞ and fcontain six components which are time-dependent functions over the domain X with the remaining fourcomponents time-dependent functions defined for on S only. The equation ðoy=osÞ þ f ¼ 0 represents 10partial differential equations with 10 unknowns and g ¼ 0 represents appropriate boundary conditions.

    2.6. Numerical solution technique

    In order to convert the system of partial differential and algebraic equations to a system of ODEs, it is

    necessary to approximate the spatial derivatives oy=on and oy=og at grid points ðni; gjÞ, i ¼ 0; . . . ;N ,j ¼ 0; . . . ;M . We choose to specify the grid in the computational domain, Fig. 2, according to the followingGauss–Lobatto Chebyshev distribution:

    ni ¼1

    2

    h1:� cos p

    Ni

    � �i; i ¼ 0; . . . ;N ;

    gj ¼1

    2

    h1:� cos p

    Mj

    � �i; j ¼ 0; . . . ;M :

    ð80Þ

    This choice of nodes is not unique and is made because global Lagrange polynomial approximations of

    general nonperiodic functions defined on this grid were found in [29] to yield a more uniform and overall

    lower error than a uniform grid. The functions yðn; g; sÞ are approximated in terms of a double Lagrangeglobal interpolating polynomial defined on the mesh nn, n ¼ 0; . . . ;N , gm, m ¼ 0; . . . ;M , i.e.

    yðn; g; sÞ �XNn¼0

    XMm¼0

    yðnn; gm; sÞLðNÞn ðnÞLðMÞm ðgÞ: ð81Þ

    The Lagrange interpolating polynomials are

    LðNÞn ðnÞ ¼QN

    l¼0; l 6¼n n� nlð ÞQNl¼0; l 6¼n nn � nlð Þ

    ; n ¼ 0; . . . ;N ;

    LðMÞm ðgÞ ¼QM

    l¼0; l 6¼m g� glð ÞQMl¼0; l 6¼m gm � glð Þ

    ; m ¼ 0; . . . ;M :

    ð82Þ

    It is easily shown that the Lagrange interpolating polynomials, LðNÞn ðnÞ and LðMÞm ðgÞ, have the values of unityat n ¼ nn and g ¼ gm and zero at the other collocation points, i.e.

    LðNÞn ðniÞ ¼ dni ¼0 if n 6¼ i;1 if n ¼ i;

    �LðMÞm ðgjÞ ¼ dmj ¼

    0 if m 6¼ j;1 if m ¼ j:

    �ð83Þ

    Derivatives of yðn; g; s; bÞ are evaluated by differentiating Eq. (81). Evaluating these derivatives on the grid,ðni; gjÞ, chosen to be the same grid as that used to define the interpolating polynomial, i.e. ðni; gjÞ � ðnn; gmÞ,and making use of Eq. (83) yields

    oy

    on

    ni ;gjð Þ

    �XNn¼0

    yðnn; gj; sÞdLndn

    ðniÞ;

    oy

    og

    ni ;gjð Þ

    �XMm¼0

    yðni; gm; sÞdLmdg

    ðgjÞ:

    ð84Þ

    The terms ðdLn=dnÞðniÞ and ðdLm=dgÞðgjÞ in Eq. (84) are evaluated efficiently for an arbitrary grid using analgorithm developed by Fornberg [25]. The operation count for approximating the derivatives via Eq. (84)

    on an N �M grid is ðNMÞ2 operations for direct matrix multiplication used here. The points which both

  • G.P. Brooks, J.M. Powers / Journal of Computational Physics 197 (2004) 58–85 73

    define the Lagrange interpolating polynomials and at which derivatives are evaluated are chosen according

    to Eq. (80). The metrics o2z=ðos onÞ and o2r=ðos onÞ in Eq. (70) are specified by differentiating Eq. (84) withrespect to time, i.e.

    o2zðni; gj; sÞos on

    ¼XNn¼0

    o

    oszðnn; gj; sÞ

    dLndn

    ðniÞ;

    o2rðni; gj; sÞos on

    ¼XNn¼0

    o

    osrðnn; gj; sÞ

    dLndn

    ðniÞ:

    ð85Þ

    After spatial discretization of Eqs. (77)–(79) on an ðN þ 1Þ � ðM þ 1Þ grid, the equations become a systemof differential algebraic equations of the form in Eqs. (5) and (6) consisting of P2 ¼ 6ðNM þ N þMÞ þ 5total equations and equal number of unknowns. The system is composed of P1 ¼ 6NM þ 2M ODEs and6N þ 4M þ 5 algebraic equations, where the primary variables, ypðsÞ, p ¼ 1; . . . ; P1, taken to be those whosetime derivative explicitly appears in Eq. (5) are

    yp sð Þ ¼

    q ni; gj; s� �

    u ni; gj; s� �

    w ni; gj; s� �

    p ni; gj; s� �

    9>>>>>=>>>>>;i ¼ 1; . . . ;N ; j ¼ 1; . . . ;M � 1;

    r ni; gj; s� �z ni; gj; s� �) i ¼ 0; . . . ;N ; j ¼ 1; . . . ;M ;vBT ni; sð Þ

    qðni; 0; sÞ

    pðni; 0; sÞ

    vðni; sÞ

    9>>>>>=>>>>>;i ¼ 1; . . . ;N ;

    26666666666666666666666664

    37777777777777777777777775

    ; p ¼ 1; . . . ; P1; ð86Þ

    and the secondary variables yp0 ðsÞ, p0 ¼ P1 þ 1; . . . ; P2; are

    yp0 ðsÞ ¼

    u ni; 0; sð Þw ni; 0; sð Þq ni; 1; sð Þu ni; 1; sð Þw ni; 1; sð Þp ni; 1; sð Þ

    9>>>>>>>>>>=>>>>>>>>>>;i ¼ 1; . . . ;N ;

    qð0; gj; sÞuð0; gj; sÞwð0; gj; sÞpð0; gj; sÞ

    9>>>>=>>>>; j ¼ 0; . . . ;M ;v 0; sð Þ

    26666666666666666666666664

    37777777777777777777777775

    ; p0 ¼ P1 þ 1; . . . ; P2: ð87Þ

  • 74 G.P. Brooks, J.M. Powers / Journal of Computational Physics 197 (2004) 58–85

    The functions fpðy1; . . . ; yP2Þ, p ¼ 1; . . . ; P1, are

    fp ¼

    �bu oqon

    � bw oqog

    �q onor

    ouon

    þonoz

    owon

    þogor

    ouog

    þogoz

    owog

    � ��qu

    r

    ni ;gj ;sð Þ

    �bu ouon

    � bw ouog

    � 1q

    onor

    opon

    þogor

    opog

    � ��

    ni ;gj ;sð Þ

    �buowon

    � bw owog

    � 1q

    onoz

    opon

    þogoz

    opog

    � ��

    ni ;gj;sð Þ

    �buopon

    � bw opog

    � cp onor

    ouon

    þonoz

    owon

    þogor

    ouog

    þogoz

    owog

    � �� cpu

    r

    ni ;gj;sð Þ

    9>>>>>>>>>>>>>>>=>>>>>>>>>>>>>>>;

    i¼ 1; . . . ;N ; j¼ 1; . . . ;M�1;

    gdZðnÞdn

    v n;sð ÞffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffidRðnÞdn

    � �2þ dZðnÞ

    dn

    � �2s

    ni ;gj ;sð Þ

    �gdRðnÞdn

    v n;sð ÞffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffidRðnÞdn

    � �2þ dZðnÞ

    dn

    � �2s

    ni ;gj ;sð Þ

    9>>>>>>>>>>>>>>>>>=>>>>>>>>>>>>>>>>>;

    i¼ 0; . . . ;N ; j¼ 1; . . . ;M ;

    �bu og

    orowon

    �ogoz

    ouon

    � �þ 1q

    ogor

    onoz

    �ogoz

    onor

    � �opon

    � ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiogoz

    � �2þ

    ogor

    � �2s

    ni ;0;sð Þ

    � 1c2

    opos

    þbu oqon

    � 1c2

    opon

    � ��

    ni ;0;sð Þ

    �qcbu og

    orouon

    þogoz

    owon

    � �þc og

    oronor

    þogoz

    onoz

    � �oponffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

    ogoz

    � �2þ og

    or

    � �2s266664

    �qc2 onor

    ouon

    þonoz

    owon

    þogor

    ouog

    þogoz

    owog

    � �

    þc

    ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiogoz

    � �2þ og

    or

    � �2sopog

    �bu opon

    �qc2ur

    #

    ni ;0;sð Þ

    �A2þqcA1ð Þ v1� veg

    � ��oeSNos

    �qc v� veg� �

    �oeSNos

    þA3eg � eN� �

    A2þqc A1�1ð Þ½ �

    264375

    ni ;1;sð Þ

    9>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>=>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>;

    i¼ 1; . . . ;N ;

    26666666666666666666666666666666666666666666666666666666666666666666666666666666666666666664

    37777777777777777777777777777777777777777777777777777777777777777777777777777777777777777775ð88Þ

  • G.P. Brooks, J.M. Powers / Journal of Computational Physics 197 (2004) 58–85 75

    and the functions gp0 ðy1; . . . ; yP2Þ, p0 ¼ P1 þ 1; . . . ; P2, are

    gp0 ¼

    u�vBT

    oronffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

    oron

    � �2þ

    ozon

    � �2s266664

    377775

    ni;0;sð Þ

    w�vBT

    ozonffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

    oron

    � �2þ oz

    on

    � �2s266664

    377775

    ni;0;sð Þ

    q ni; 1; sð Þ �d1 ni; sð ÞdS ni; sð Þ

    q1

    v1 � eST � v � eST½ �j ni ;1;sð Þ

    dS ni; sð Þ �c� 1cþ 1 d1 ni; sð Þ þ

    2cp1cþ 1ð Þq1d1 ni; sð Þ

    p ni; 1; sð Þ �2q1cþ 1 d

    21 ni; sð Þ �

    c� 1cþ 1 p1

    9>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>=>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>;

    i ¼ 1; . . . ;N ;

    q 0; gj; s� �

    � q 0; 1; sð Þp 0; gj; s� �p 0; 1; sð Þ

    � �1=cu 0; gj; s� �owon

    0;gj;sð Þ

    opon

    0;gj;sð Þ

    9>>>>>>>>>>>>>=>>>>>>>>>>>>>;j ¼ 0; . . . ;M ;

    o

    onv 0; sð Þ

    26666666666666666666666666666666666666666666666666666666666664

    37777777777777777777777777777777777777777777777777777777777775

    : ð89Þ

    There are no equations for the grid points on the body since these are fixed in time, nor is there an equation

    for the tangential velocity on the body at the centerline since this quantity is redundant, the velocity

    components, uð0; gj; sÞ, wð0; gj; sÞ, already being specified by Eqs. (72) and (74).Finally, the secondary variables yp0 , p0 ¼ P1 þ 1; . . . ; P2, in Eq. (89) are solved for explicitly in terms

    of the primary variables yp, p ¼ 1; . . . ; P1. Making use of Eq. (84) yields the following expression forwð0; gj; sÞ,

    w 0; gj; s� �

    ¼PN

    n¼1 wðnn; gj; sÞ dLndn ð0ÞdL0dn ð0Þ

    ; ð90Þ

    similar expressions are found for pð0; gj; sÞ, and vð0; sÞ. Eqs. (53) and (54) are reformulated into the fol-lowing two equations for the quantities uðni; 1; sÞ and wðni; 1; sÞ, i ¼ 1; . . . ;N ,

  • 76 G.P. Brooks, J.M. Powers / Journal of Computational Physics 197 (2004) 58–85

    u ni; 1; sð Þ ¼ozon

    oron

    ffiffiffic

    pM1

    ozon

    � �2þ or

    on

    � �2264 þ ozon dþ eg � eSN� �v� ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

    ozon

    � �2þ or

    on

    � �2r3775

    ni ;1;sð Þ

    ;

    w ni; 1; sð Þ ¼ozon

    � �2 ffiffiffic

    pM1

    ozon

    � �2þ or

    on

    � �2264 � oron dþ eg � eSN� �v� ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

    ozon

    � �2þ or

    on

    � �2r3775

    ni;1;sð Þ

    :

    ð91Þ

    Once the quantities wð0; gj; sÞ, pð0; gj; sÞ, zð0; gj; sÞ, and vð0; sÞ are found from Eq. (90) and uðni; 1; sÞ andwðni; 1; sÞ are found from Eq. (91), the algebraic equations, gp0 ðy1; . . . ; yP2Þ, Eq. (89), are written in the formof Eq. (8) where

    bgp0 ¼

    vBToronffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

    oron

    � �2þ oz

    on

    � �2s266664

    377775

    ni ;0;sð Þ

    vBTozonffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

    oron

    � �2þ oz

    on

    � �2s266664

    377775

    ni ;0;sð Þ

    d1 ni; sð ÞdS ni; sð Þ

    q1

    ozon

    oron

    ffiffiffic

    pM1

    ozon

    � �2þ or

    on

    � �2 þozon

    dS þ eg � eSN� �

    v� ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiozon

    � �2þ or

    on

    � �2s266664

    377775

    ni ;1;sð Þ

    ozon

    � �2 ffiffiffic

    pM1

    ozon

    � �2þ

    oron

    � �2 �oron

    dS þ eg � eSN� �

    v� ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiozon

    � �2þ

    oron

    � �2s266664

    377775

    ni ;1;sð Þ

    2q1cþ 1 d

    21 ni; sð Þ þ

    c� 1cþ 1 p1

    9>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>=>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>>;

    i ¼ 1; . . . ;N ;

    q 0; 1; sð Þ

    PNn¼1 pðnn; gj; sÞ

    dLndn

    ð0Þ

    p 0; 1; sð Þ dL0dn

    ð0Þ

    0BB@1CCA

    1=c26643775

    0;gj ;sð Þ0PN

    n¼1 wðnn; gj; sÞdLndn

    ð0Þ

    dL0dn

    ð0ÞXNn¼1

    pðnn; gj; sÞdLndn

    ð0Þ

    dL0dn

    ð0Þ

    9>>>>>>>>>>>>>>>>>>>>>=>>>>>>>>>>>>>>>>>>>>>;

    j ¼ 0; . . . ;M ;

    PNn¼1

    vðnn; sÞdLndn

    ð0ÞdL0dn

    ð0Þ

    2666666666666666666666666666666666666666666666666666666666666666666666666666664

    3777777777777777777777777777777777777777777777777777777777777777777777777777775

    ; ð92Þ

  • G.P. Brooks, J.M. Powers / Journal of Computational Physics 197 (2004) 58–85 77

    and dSðni; sÞ ¼ ððc� 1Þ=ðcþ 1ÞÞd1ðni; sÞ � 2cp1=ððcþ 1Þq1d1ðni; sÞÞ.The initial conditions for the shock distance, hðni; sÞ, are taken to be constant, i.e., hðni; sÞ ¼

    hð0; sÞ ¼ 0:25, i ¼ 1; . . . ;N , which is sufficient to set initial conditions for the remainder of the physical gridcoordinates, rðni; gj; sÞ, zðni; gj; sÞ, i ¼ 0; . . . ;N , j ¼ 0; . . . ;M , by making use of Eqs. (34) and (35). Theshock velocity, vðni; sÞ, i ¼ 0; . . . ;N , is initially set to zero. The initial values for the variables, qðni; gj; sÞ,uðni; gj; sÞ, wðni; gj; sÞ, pðni; gj; sÞ, i ¼ 1; . . . ;N , j ¼ 1; . . . ;M , qðni; 0; sÞ, uðni; 0; sÞ, and pðni; 0; sÞ,i ¼ 1; . . . ;N , are set equal to the value behind the shock at ðg ¼ 1Þ for the corresponding n coordinate line,e.g. qðni; gj; 0Þ ¼ qðni; 1; 0Þ. The initial values for uðni; 0; sÞ, i ¼ 1; . . . ;N , are chosen so that the boundarycondition at g ¼ 0, Eq. (41) is satisfied exactly, and the initial values for qð0; gj; sÞ, uð0; gj; sÞ, wð0; gj; sÞ,pð0; gj; sÞ, j ¼ 0; . . . ;M , are chosen so that the boundary conditions at n ¼ 0, Eqs. (72)–(75) are satisfiedexactly. The initial values for the variables vBTðni; sÞ, i ¼ 1; . . . ;N , are prescribed once the values forrðni; 0; sÞ, zðni; 0; sÞ, uðni; 0; sÞ and wðni; 0; sÞ have been specified.

    Solutions have been obtained for the system of ODEs, Eqs. (9), with the standard ODE solver LSODA,

    [30,31], which automatically adjusts the time step to achieve a specified level of accuracy. It also auto-

    matically switches between an explicit method and implicit method depending on the stiffness of the

    problem. A typical steady state calculation on a 17� 9 grid took 106 s CPU time on a single 800 MHzprocessor with 512 MB of RAM. For the steady state problems, the criteria for stopping the integration is

    when the L1½X� error in qðn; g; s ! 1Þ does not change appreciably. Finally, since LSODA automaticallyswitches between an explicit method and implicit method depending, it may be of interest to note that earlyin the calculation, less than 10% of LSODA�s time steps were implicit whereas about 50% of the time stepswere run in implicit mode as the calculations neared steady state. Apparently as the level of error ap-

    proached the level of the discretization error, the problem became increasingly stiff causing LSODA to

    switch to implicit mode.

    2.7. Pseudospectral flow solver verification and validation

    2.7.1. Taylor–Maccoll solution

    The solution to supersonic flow over a cone, also known as the Taylor–Maccoll solution [32], will be

    used to verify the code described in the previous section. We first use a highly accurate ODE solver to

    calculate the Taylor–Maccoll solution, which we will subsequently refer to as the exact solution. The only

    modification to the blunt body problem formulation to generate numerical approximations to the Taylor–

    Maccoll flow is to replace the centerline boundary condition at n ¼ 0 with a Dirichlet boundary conditioncontaining the values of q, u, w, p, r, and z taken from the exact solution. A schematic of a 40� cone in-cluding the physical and computational coordinates is shown in Fig. 3 for a 5� 5 grid for M1 ¼ 3:5,q1 ¼ p1 ¼ 1. A value of r0 ¼ 0:1 was chosen for the results presented in this section. The initial conditionsfor q, u, w, p, r, and z are taken from the exact solution, and a sinusoidal distribution is chosen for the initialshock velocity, i.e. vðn; 0Þ ¼ 0:1 sinð2pnÞ. In Fig. 4, we show the time history of the L1½X� error in qðn; gÞover the domain, X, for the pseudospectral prediction measured against the exact solution for a M1 ¼ 3:5flow over a 40� cone solved on a 5� 17 grid. The figure demonstrates a rapid relaxation to the exactsolution.

    A grid convergence test for the pseudospectral prediction of the Taylor–Maccoll flow is conducted by

    refining the grid in the g-direction for a fixed number of 5 nodes in the n-direction. The accuracy of themethod is unaffected by grid refinement in the n-direction since all derivatives are zero in that direction. Aswe can see from Fig. 5, there is a rapid decrease in the error until about 10�12 when the error flattens

    probably due to roundoff error. Note the spectral nature of the grid convergence, that is the slope of the

    error curve continues to steepen with increasing number of nodes, at least until the roundoff limit is

    reached, and does not reach a constant value for the slope as would be the case for a method with a fixed

    order of accuracy.

  • Fig. 3. Schematic of the physical ðr; zÞ and computational ðn; gÞ grids for the Taylor–Maccoll problem.

    2 4 6 8 1010

    15

    1010

    105

    100

    τ

    L ∞[Ω

    ] err

    or in

    ρ

    Fig. 4. Single 5� 17 grid L1½X� residual error in qðn; gÞ measured against Taylor–Maccoll similarity solution as a function of time, s,for a 40� cone at M1 ¼ 3:5.

    78 G.P. Brooks, J.M. Powers / Journal of Computational Physics 197 (2004) 58–85

    2.7.2. Steady state blunt body results

    The following functions have been chosen to parameterize the blunt body surface

    RðnÞ ¼ n; ð93Þ

    ZðnÞ ¼ n1=b; ð94Þ

    where the domain for the geometric parameter b is restricted to b 2 ð0; 1Þ. Eliminating the parameter n, wesee that the body surface is described by R ¼ Zb. For b ¼ 0:5, M1 ¼ 3:5, q1 ¼ p1 ¼ 1, contour plots ofMach number and pressure are shown in Figs. 6 and 7. The sonic line, M ¼ 1, is predicted in Fig. 6 as wellas the fact that the outflow velocity is indeed supersonic as required in the derivation of the outflow

  • 100

    101

    102

    103

    1015

    1010

    105

    100

    ηnumber of nodes in direction

    L ∞[Ω

    ] err

    or in

    ρ

    Fig. 5. L1½X� error in qðn; gÞ measured against a Taylor–Maccoll similarity solution for a 40� cone in M1 ¼ 3:5 flow as grid is refinedin the g direction.

    0.2 0 0.2 0.4 0.6 0.80

    0.5

    1

    1.5

    z

    r

    0.2

    0.4

    0.6

    0.8

    1

    1.2

    1.4

    1.4

    1.6

    1.6

    1.81.8

    Fig. 6. Contours of Mach number for flow over the blunt body for b ¼ 0:5, M1 ¼ 3:5, 17� 9 grid.

    G.P. Brooks, J.M. Powers / Journal of Computational Physics 197 (2004) 58–85 79

    boundary condition. In Fig. 7, the pressure at the stagnation point is seen to be more than 16 times the

    freestream pressure at M1 ¼ 3:5, and the jump in pressure across the normal shock at the centerline is over13 times the freestream pressure.

    As a means of code validation, a comparison is made between the numerical results for the pressure

    distribution on the body with that of the modified Newtonian [33] sine squared law

    Cp ¼ Cp0 sin2 /; ð95Þ

    where Cp0 is the pressure coefficient at the body stagnation point and / is the local surface inclination anglemeasured with respect to the z axis. The pressure coefficient, Cp, is defined as

  • 0.2 0 0.2 0.4 0.6 0.80

    0.5

    1

    1.5

    z

    r

    5.5

    6.5

    7.5

    8.5

    10

    11.5

    13.5

    1516

    Fig. 7. Contours of pressure for flow over the blunt body for b ¼ 0:5, M1 ¼ 3:5, 17� 9 grid.

    80 G.P. Brooks, J.M. Powers / Journal of Computational Physics 197 (2004) 58–85

    Cp ¼p� � p�112q�1w�21

    ¼ p n; 0; sð Þ � 112cM21

    : ð96Þ

    The modified Newtonian approximation is a semi-analytical model for the surface pressure distribution

    over blunt bodies. Anderson [34] reports that for a power law body with b ¼ 0:5 and aspect ratio near unity,the modified Newtonian approximation does well in predicting the pressure distribution on the surface of

    the body. As can be seen from Fig. 8, the pseudospectral code also predicts close agreement for the pressure

    0 0.2 0.4 0.6 0.8 10.2

    0.4

    0.6

    0.8

    1

    1.2

    1.4

    1.6

    1.8

    r

    Cp

    Pseudospectral predictionModified Newtonian theory

    Fig. 8. Blunt body surface Cp distribution predictions at M1 ¼ 3:5 from modified Newtonian theory and from the pseudospectralmethod, where b ¼ 0:5; 17� 9 grid.

  • G.P. Brooks, J.M. Powers / Journal of Computational Physics 197 (2004) 58–85 81

    distribution on the surface of the body defined by r ¼ffiffiz

    p. As a further check on the validity of the

    pseudospectral code in Fig. 9 a comparison is made of the pseudospectral prediction for the shock shape for

    M1 ¼ 3:5 flow over a sphere with that of an empirical formula by Billig [35] developed for flow overspherically blunted cones based on experiment.

    A grid convergence study is performed for the blunt body with the L1½X� error over the domain, X inqðn; gÞ shown in Fig. 10, at M1 ¼ 3:5 and b ¼ 0:5, where the error is measured against a 65� 33 or 2145node numerical solution. For 861 nodes, the L1½X� error over the domain, X in qðn; gÞ, has been reduced tothe order of 10�12 and subsequently flattens due to roundoff error. Like grid convergence plots for the

    Taylor–Maccoll solution, the convergence of the error for the blunt body problem shows a spectral con-

    vergence rate as expected of the pseudospectral numerical technique.

    0.6 0.4 0.2 0 0.2 0.4 0.6 0.8 10

    0.2

    0.4

    0.6

    0.8

    1

    1.2

    1.4

    z

    r

    Body surfacePseudospectral predictionBillig

    Fig. 9. Shock shape prediction of the pseudospectral code for a sphere M1 ¼ 3:5 compared with an empirical formula by Billig [35]derived from experiments; 17� 9 grid.

    101

    102

    103

    1010

    105

    number of nodes

    L ∞[Ω

    ] err

    or in

    ρ

    Fig. 10. Grid convergence L1½X� error in qðn; gÞ measured against a baseline, 65� 33 grid, solution for a b ¼ 0:5, M1 ¼ 3:5 bluntbody.

  • 82 G.P. Brooks, J.M. Powers / Journal of Computational Physics 197 (2004) 58–85

    2.7.3. Unsteady acoustic wave interaction with attached shock

    Now we turn to a fundamentally unsteady problem: the interaction of an unsteady, planar acoustic wave

    with an attached shock, for which the automatic time step selection in LSODA enables us to have tightcontrol over the error in the solution. A schematic of the grid is shown in Fig. 3, where r0 ¼ 0:01 is chosenfor this problem. The values of the freestream flow quantities q1, u1, w1, p1 at the shock are taken to be

    q1 ¼ �1

    2ffiffiffic

    p F n; sð Þð � G n; sð ÞÞ; ð97Þ

    u1 ¼ 0; ð98Þ

    w1 ¼1

    2F n; sð Þð þ G n; sð ÞÞ; ð99Þ

    p1 ¼ 1þ c q1ð � 1Þ; ð100Þwhere

    F n; sð Þ ¼ ffiffifficp ðM1 � 1Þ; ð101Þ

    G n; sð Þ ¼ ffiffifficp ðM1 þ 1Þð1þ � sin jðzðn; 1; sÞ � ffiffifficp ðM1 þ 1ÞsÞÞ: ð102Þ

    Choosing � ¼ 0:01, j ¼ 6p, M1 ¼ 3:5, and c ¼ 1:4 in Eqs. (101) and (102), the system of ODEs in Eq. (9)are integrated in time for s 2 ½0; 2� on a 33� 17 grid using the ODE solver LSODA. The CPU time for thecalculation on a single 800 MHz processor was 7.5 h. The initial conditions were set to the Taylor–Maccoll

    solution of the unperturbed freestream flow conditions, i.e. q1 ¼ 1, u1 ¼ 0, w1 ¼ffiffiffic

    pM1, p1 ¼ 1. The

    relative and absolute error tolerances for the time integration were set to 10�10 and 10�12, respectively, in

    LSODA. As an estimate for the spatial accuracy, we recall that the L1½X� error in qðn; gÞ on a 33� 17 gridfor the steady state calculations was 10�12 and 10�9 for the Taylor–Maccoll and blunt body flows,

    respectively.

    Let us analyze the motion of a single point on the shock located at n ¼ g ¼ 1. The power spectrum, Pðf Þ,as a function of reduced frequency for the perturbation in freestream density, Dq1 ¼ q1 � 1, and shockdistance function, Dh ¼ hð1; tÞ � h1ð1Þ, at the point n ¼ g ¼ 1 are presented in Fig. 11 as well as Dq1jz¼�1,which is well upstream of the shock. Initial transients in the solution are neglected in the estimation of the

    power spectrum by considering only s 2 ½1:001; 2�, so that the time interval is T ¼ 0:999. The powerspectrum is defined at K=2þ 1 frequencies as

    P ðf0Þ ¼1

    K2jC0j2;

    P ðfkÞ ¼1

    K2jCkj2h

    þ jCK�kj2i; k ¼ 1; . . . ; K

    2

    �� 1

    �;

    P ðfK=2Þ ¼1

    K2jCK=2j2;

    ð103Þ

    where fk is defined only for the zero and positive frequencies

    fk ¼kKD

    ; k ¼ 0; . . . ; K2

    � �; ð104Þ

    where D ¼ 0:001 is the chosen sampling interval, K ¼ T=Dþ 1 ¼ 1000 is the number of sampled points, andthe Ck are the discrete Fourier coefficients. The power spectrum, Fig. 11, clearly shows a large peak in the

  • 0 20 40 60 80 10010

    12

    1010

    108

    106

    104

    102

    100

    reduced frequency (fk)

    P(f

    k)

    ∆ ρ∞|z= 1∆ ρ∞∆ h

    Fig. 11. Frequency spectrum at a single point on the shock ðn ¼ 1Þ of the fluctuations in freestream density, Dq1 ¼ q1 � 1, and theresponse of the shock, Dh ¼ hð1; tÞ � h1ð1Þ. M1 ¼ 3:5; 33� 17 grid.

    G.P. Brooks, J.M. Powers / Journal of Computational Physics 197 (2004) 58–85 83

    power spectrum for Dq1jz¼�1 at a reduced frequency of f16 ¼ ðj=2pÞffiffiffic

    p ðM1 þ 1Þ ¼ 16:0, according to Eq.(9). Peaks in the power spectrum of Dh, and Dq1 also appear at f16 ¼ 16:0 in response to the frequency ofoscillation of Dq1jz¼�1. In addition, higher harmonics at integer multiples of the f16 ¼ 16:0 �forcing� fre-quency are present in the power spectrum of Dh and Dq1.

    The time history of the density fluctuations and perturbation in shock distance function both measured

    at the point n ¼ g ¼ 1 are shown in Fig. 12 for s 2 ½1:75; 2:0�. It is evident from Fig. 12 that the freestreamdensity fluctuations are nearly 180� out of phase with the shock distance perturbations, and that the am-plitude of the density fluctuations are 20% of the mean flow, while the shock response is two orders of

    magnitude below the perturbations in freestream density. The high accuracy of the current method is

    critical in capturing the correct shock dynamics for such small amplitude fluctuations, and in predicting the

    higher harmonics of the shock fluctuations in Fig. 11, where the amplitude of the power spectrum drops

    1.7 1.75 1.8 1.85 1.9 1.95 2-0.3

    -0.2

    -0.1

    0

    0.1

    0.2

    0.3

    ∆ρ∞

    1.7 1.75 1.8 1.85 1.9 1.95 2-6

    -4

    -2

    0

    2

    4

    6 x 10-3

    ∆h∞

    τ

    ∆ ρ∞

    ∆ h

    Fig. 12. Time history at a single point on the shock ðn ¼ 1Þ of the fluctuations in freestream density, Dq1 ¼ q1 � 1, and the responseof the shock, Dh ¼ hð1; tÞ � h1ð1Þ. M1 ¼ 3:5; 33� 17 grid.

  • 84 G.P. Brooks, J.M. Powers / Journal of Computational Physics 197 (2004) 58–85

    several orders of magnitude with each successively increasing harmonic. The current method is able to

    resolve up to the fourth harmonic of Dh, whose amplitude in the power spectrum is on the order of 10�10.

    3. Conclusion

    In this study, we have described a pseudospectral numerical approximation technique for the inviscid

    supersonic flow over a blunt body geometry in which the discretized form of the governing equations and

    boundary conditions are formulated in terms of a system of ODE�s which can be solved using a standardODE solver. This formulation leverages the strengths of widely available ODE solvers to generate time

    accurate solutions within prescribed error tolerances through automatic time step selection and explicit/implicit switching. Additionally, fitting of the shock and the use of global polynomials in the solution

    approximation permit high accuracy steady state approximations with 10�5 error in as little as 106 s on a

    single 800 MHz processor. Even higher accuracy has been achieved for a time-dependent problem. This

    standard formulation has important potential applications such as approximating unsteady shock phe-

    nomena with sufficient accuracy to discern between physical versus numerical instabilities.

    References

    [1] G.P. Brooks, J.M. Powers, A Karhunen-Lo�eve least-squares technique for optimization of geometry of a blunt body in supersonic

    flow, J. Comput. Phys. (2004), in press.

    [2] D. Gottlieb, S. Orszag, Numerical Analysis of Spectral Methods: Theory and Applications, SIAM-CBMS, Philadelphia, 1977.

    [3] N.K Yamaleev, M.H. Carpenter, On accuracy of adaptive grid methods for captured shocks, J. Comput. Phys. 181 (2002) 280–

    316.

    [4] S.P. D�yakov, On the stability of shock waves, ZETF 27 (1954) 288–295.[5] V.M. Kontorovich, Concerning the stability of shock waves, Sov. Phys. JETP 6 (1957) 1179–1180.

    [6] J.W. Bates, D.C. Montgomery, The D�yakov instability of shock waves in real gases, Phys. Rev. Lett. 84 (6) (2000) 1180–1183.[7] K.M. Peery, S.T. Imlay, Blunt body flow simulations, AIAA Paper 88-2924, 1988.

    [8] J.-Ch. Robinet, J. Gressier, G. Casalis, J.-M. Moschetta, Shock wave instability and the carbuncle phenomenon: same intrinsic

    origin?, J. Fluid Mech. 417 (2000) 237–263.

    [9] J.J. Quirk, A contribution to the great Riemann solver debate, Int. J. Numer. Meth. Fluids 18 (1994) 555.

    [10] J. Gressier, J.-M. Moschetta, Robustness versus accuracy in shock-wave computations, Int. J. Numer. Meth. Fluids 33 (2000)

    313–332.

    [11] J.-F. Coulombel, S. Benzoni-Gavage, D. Serre, Note on a Paper by Robinet, Gressier, Casalis and Moschetta, J. Fluid Mech. 469

    (2002) 401–405.

    [12] V.V. Rusanov, A blunt body in a supersonic stream, Annu. Rev. Fluid Mech. 8 (1976) 377–404.

    [13] W.D. Hayes, R.F. Probstein, Hypersonic flow theory, Inviscid Flows, vol. 1, Academic Press, New York, 1966.

    [14] C.C. Lin, S.I. Rubinov, On the flow behind curved shocks, J. Math. Phys. 27 (1948) 105–129.

    [15] P.R. Garabedian, H.M. Lieberstein, On the numerical calculation of detached bow shock waves in hypersonic flow, J. Aero Sci. 25

    (1) (1958) 109–118.

    [16] M.D. Van Dyke, The supersonic blunt-body problem – review and extension, J. Aero/Space Sci. 25 (4) (1958) 485–496.

    [17] M.W. Evans, F.H. Harlow, Calculation of supersonic flow past an axially symmetric cylinder, J. Aero Sci. 25 (1) (1958) 269.

    [18] G. Moretti, M. Abbett, A time-dependent computational method for blunt body flows, AIAA J. 4 (12) (1966) 2136–2141.

    [19] M.Y. Hussaini, D.A. Kopriva, M.D. Salas, T.A. Zang, Spectral methods for the Euler equations: Part 2. Chebyshev methods and

    shock-fitting, AIAA J. 23 (2) (1985) 234–240.

    [20] D.A. Kopriva, T.A. Zang, M.Y. Hussaini, Spectral methods for the Euler equations: the blunt body problem revisited, AIAA J. 29

    (9) (1991) 1458–1462.

    [21] D.A. Kopriva, Shock-fitted multidomain solution of supersonic flows, Comput. Meth. Appl. Mech. Eng. 175 (1999) 383–394.

    [22] G.P. Brooks, J.M. Powers, A Karhunen–Lo�eve Galerkin technique with shock-fitting for optimization of a blunt body geometry,

    AIAA Paper 2002-3861, 2002.

    [23] C. Canuto, M.Y. Hussaini, A. Quarteroni, T.A. Zang, Spectral Methods in Fluid Dynamics, Springer, New York, 1988.

    [24] D. Gottlieb, J.S. Hesthaven, Spectral methods for hyperbolic problems, J. Comput. Appl. Math. 128 (2001) 83–131.

    [25] B. Fornberg, A Practical Guide to Psuedospectral Methods, Cambridge University Press, Cambridge, 1998.

  • G.P. Brooks, J.M. Powers / Journal of Computational Physics 197 (2004) 58–85 85

    [26] B.A. Finlayson, The Method of Weighted Residuals and Variational Principles, Academic Press, New York, 1972.

    [27] K.W. Thompson, Time dependent boundary conditions for hyperbolic systems, J. Comput. Phys. 68 (1987) 1–24.

    [28] K.W. Thompson, Time dependent boundary conditions for hyperbolic systems, II, J. Comput. Phys. 89 (1990) 439–461.

    [29] G.P. Brooks, A Karhunen–Lo�eve least-squares technique for Optimization of Geometry of a Blunt Body in Supersonic Flow,

    Ph.D. Dissertation, University of Notre Dame, Notre Dame, IN, 2003.

    [30] A.C. Hindmarsh, ODEpack, A systematized collection of ODE solvers, in: R.S. Stepleman et al. (Eds.), Scientific Computing,

    North-Holland, Amsterdam, 1983, p. 55.

    [31] L.R. Petzold, Automatic selection of methods for solving stiff and nonstiff systems of ordinary differential equations, SIAM J. Sci.

    Comput. 4 (1983) 136.

    [32] G.I. Taylor, J.W. Maccoll, The air pressure on a cone moving at high speeds, Proc. R. Soc. A 139 (1933) 278–311.

    [33] L. Lees, Hypersonic flow, in: Proceedings of the Fifth International Aeronautical Conference, Los Angeles, Institute of the

    Aeronautical Sciences, New York, 1955, pp. 241–276.

    [34] J.D. Anderson, Hypersonic and High Temperature Gas Dynamics, McGraw-Hill, New York, 1989.

    [35] F.S. Billig, Shock-wave shapes around spherical and cylinder-nosed bodies, J. Spacecraft Rockets 4 (6) (1967) 822–823.

    Standardized pseudospectral formulation of the inviscid supsersonic blunt body problemIntroductionSupersonic blunt body flow and pseudospectral solverStandard formulationGoverning equationsComputational and physical coordinatesBoundary conditionsSummary of governing equations and boundary conditionsNumerical solution techniquePseudospectral flow solver verification and validationTaylor-Maccoll solutionSteady state blunt body resultsUnsteady acoustic wave interaction with attached shock

    ConclusionReferences